Sunteți pe pagina 1din 13

Overview

When used to study electrochemical systems, electro-


chemical impedance spectroscopy (EIS) can give you
accurate, error-free kinetic and mechanistic information
using a variety of techniques and output formats. For
this reason, EIS is becoming a powerful tool in the
study of corrosion, semiconductors, batteries, electro-
plating, and electro-organic synthesis.
Table I summarizes some of the electrochemical phe-
nomena that have been studied using EIS. In these
areas, EIS offers three advantages over dc techniques:
Small Amplitude: EIS techniques use very small exci-
tation amplitudes, often in the range of 5 to 1 0 mV
peak-to-peak. Excitation waveforms of this amplitude
cause only minimal perturbation of the electrochemical
test system, reducing errors caused by the measure-
ment technique.
Research Area Application
Corrosion
(See references 1-15)
Rate Determinations
Inhibitor and Coatings
Passive Layer Investiga-
tions
Coatings Evaluation
(See references 16-21)
Dielectric Measurements
Corrosion Protection
Batteries
(See references 22-28)
State-of-charge
Materials Selection
Electrode Design
Electrodeposition
(See references 29-32)
Bath Formulation
Surface Pretreatment
Deposition Mechanism
Deposit Characterization
Electro-Organic Synthesis
(see reference 33)
Adsorption/Desorption
Reaction Mechanism
Semiconductors
(See references 34-36)
Photovoltaic work
Dopant Distributions
Table 1: Applications Amenable to Impedance Studies
Mechanism Study: Because electrochemical imped-
ance experiments provide data on both electrode ca-
pacitance and charge-transfer kinetics, EIS techniques
can provide valuable mechanistic information.
Measurement Accuracy: Because the method does
not involve a potential scan, you can make measure-
ments in low conductivity solutions where dc techniques
are subject to serious potential-control errors. In fact,
you can use EIS to determine the uncompensated re-
sistance of an electrochemical cell.
Theoretical Advantages
The main advantage of EIS is that you can use a purely
electronic model to represent an electrochemical cell.
An electrode interface undergoing an electrochemical
reaction is typically analogous to an electronic circuit
consisting of a specific combination of resistors and
capacitors. You can take advantage of this analogy by
using established ac circuit theory to characterize the
electrochemical system in terms of its equivalent circuit.
In practice, you can correlate an impedance plot ob-
tained for a given electrochemical system w4th one or
more equivalent circuits. You can use this information
to verify a mechanistic model for the system, or at least
to rule out incorrect models. Once you choose a par-
ticular model, you can correlate physical or chemical
properties with circuit elements and extract numerical
values by fitting the date to the circuit model.
About This Note
Because EIS generates detailed information, sophisti-
cated approaches are required to interpret the data and
extract meaningful results. The recent increase in litera-
ture demonstrates that a better understanding of im-
pedance theory and data interpretation has followed the
advances in instrumentation. Two of the better longer
works are Digby D. Macdonald's 'Transient Techniques
in Electrochemistry" and J. Ross Macdonalds 'Imped-
ance Spectroscopy.
37,38
In this note, we will review basic EIS theory and de-
scribe the commonly used plot formats. We will also
introduce some of the simpler methods of data interpre-
tation. This should give you a good starting point for
further study.
Introduction

Application Note AC-1
Subject: Basics of Electrochemical Impedance Spectroscopy

Overview
Electrochemical impedance theory is a well-developed
branch of ac theory that describes the response of a
circuit to an alternating current or voltage as a function
of frequency. The mathematics of the theory is beyond
the scope of this discussion, but we will present the
basic theory here.
In dc theory (a special case of ac theory where the frequency
equals 0 Hz) resistance is defined by Ohm's Law:
E = I R (1)
Using Ohm's law, you can apply a dc potential (E) to a
circuit, measure the resulting cur-rent (1), and compute
the resistance (R) - or determine any term of the equa-
tion if the other two are known. Potential values are
measured in volts (V), current in amperes or amps (A),
and resistance in ohms (). A resistor is the only ele-
ment that impedes the flow of electrons in a dc circuit.
In ac theory, where the frequency is non-zero, the
analogous equation is:
E = I Z (2)
As in Equation 1, E and I are here defined as potential
and current, respectively. Z is defined as impedance,
the ac equivalent of resistance. Impedance values are
also measured in ohms (). In addition to resistors, ca-
pacitors and inductors impede the flow of electrons in
ac circuits.
In an electrochemical cell, slow electrode kinetics, slow
preceding chemical reactions, and diffusion can all im-
pede electron flow, and can be considered analogous
to the resistors, capacitors, and inductors that impede
the flow of electrons in an ac circuit.
Figure I shows a typical plot of a voltage sine wave (E)
applied across a given circuit and the resultant ac cur-
rent waveform (1). Note that the two traces are different
not only in amplitude, but are also shifted in time.

with respect to each other - that is, they are out of phase. In
the case of a purely resistive network, the two waveforms
would not be shifted. They would be exactly in phase, differ-
ing only in amplitude.
The current sine wave can be described by the equation:
l(t) = A sin (wt + 0) (3)
where
l(t) = instantaneous current
A = maximum amplitude
= frequency in radians per second = 2f
( where f = frequency in Hertz)
t = time
= phase shift in radians
Vector Analysis
Vector analysis provides a convenient method of char-
acterizing an ac waveform It lets you describe the wave
in terms of its amplitude and phase characteristics.
Theory
What is Impedance?

The terms resistance and impedance both denote an opposition to
the flow of electrons or current. In direct current 9dc) circuits, only
resistors produce this effect. However, in alternation current (ac)
circuits, two other circuit elements, capacitors and inductors, im-
pede the flow of electrons. Impedance can be expressed as a
complex number, where the resistance is the real component and
the combined capacitance and inductance is the imaginary com-
ponent.
The total impedance in a circuit is the combined opposition of all
its resistors, capacitors, and inductors to the flow of electrons. The
opposition of capacitors and inductors is given the same name
reactance, symbolized by X and measured in ohms (). Since the
symbol for capacitance is C, capacitive reactance is symbolized by
XC. Similarly, since the symbol for inductance is L, inductive reac-
tance is symbolized by XL.
Capacitors and inductors affect not only the magnitude of an alter-
nating current but also its time-dependent characteristics or
phase. When most of the opposition to current flow comes from its
capacitive reactance, a circuit is said to be largely capacitive and
the current leads the applied voltage in phase angle. When most
of the opposition to current flow comes from its inductive reac-
tance, a circuit is said to be largely inductive and the current lags
the applied voltage in phase angle. The more inductive a circuit is,
the closer the difference in phase angle approaches 90 degrees.
Its sometimes easier to perform calculations using admittance, the
reciprocal of impedance. Admittance is symbolized by Y and
measured in siemens (S). Like impedance, admittance cab be
expressed as a complex number, where the conductance, the
reciproc al of resistance, is the real component, and the suscep-
tance, the reciprocal of reactance, is the imaginary component.
Figure 1: AC Waveforms for an Applied Potential and a Resulting
Current.
I
E
Time

For example, Figures 2, 3, and 4 show vector analyses
for the resultant current waveform of Figure 1. The cur-
rent waveform vector can be graphically described in a
variety of ways.
Figure 2 shows how the end point of the vector can be
described in terms of an (x, y) coordinate pair formed
from the in-phase (x) and out-of-phase (y) components.
In Figure 3, the vector is unambiguously defined by
phase angle () and current magnitude ( |I| ).
Figure 4 shows a third approach, often more convenient
for numerical analysis. The axes are defined as real (I')
and imaginary (I). The real and imaginary components
can be handled as a single number in complicated
equations if the complex number notation is used. The
mathematical convention for expressing quantities in
this coordinate system is to multiply the 1', or imagi-
nary, coordinate value by v-T.
Using this complex number convention, an ac current
vector can be defined as the sum of its real and imagi-
nary components:
I
Total
= I + I j (4)
where
j = -1
Note: Although mathematicians use i to stand
for -1, electrochemists use j, to avoid confusion with
i, the symbol for current.
Note that the location of the point in Figures 2, 3, and 4
has not changed - only the labels on the axes are dif-
ferent. The three different ways of expressing the posi -
tion of the point - (x, y) pair, phase angle/magnitude,
and real/imaginary coordinates, are really all the same.
The real and imaginary components of an ac current or
voltage waveform are defined with respect to some ref-
erence waveform. The real component is in phase with
the reference waveform, and the imaginary component,
(or quadrature component) is exactly 90 degrees out of
phase. The reference waveform allows us to express
the current and voltage waveforms as vectors with re-
spect to the same coordinate axes. This facilitates
mathematical manipulation of these vector quantities.
Specifically, this allows us to use Equation 2 to calcu-
late the impedance vector as the quotient of the voltage
and current vectors:
Z
Total
= E' + E"j
I + I"j (5)

where the ac voltage vector, E, can also be expressed as a
complex number.
ETotal = E' + E" j (6)
The resulting vector expression for the ac impedance,
ZTotal = Z' + Z" j (7)
is defined in terms of the same coordinate axes as the current
and voltage vectors.
By analogy \with Figures 3 and 4, the absolute magnitude of
the impedance (that is, the length of the vector) can be ex-
pressed as
I Z I = Z
2
+ Z
2
(8)
and the phase angle can be defined by:

tan = Z / Z (9)

Figure 2: Vector in Terms of X and Y Coordinates
Figure 3: Vector in terms of Angle () and Magnitude (|Z|)
Figure 4: Vector in Terms of Real (I') and Imaginary (I") Coordi-
nates
( X , Y )
Y
X
|I|

Real
I
m
a
g
i
n
a
r
y

I = I + jI I
I
Equivalent Circuit Elements
Given this basic theory, we can now look at impedance
expressions for some simple electrical circuits. Table 2
shows that the impedance of a resistor has no imagi-
nary component at all. The phase shift is zero degrees -
that is, the current is in phase with the voltage. Both
current and impedance are independent of the fre-
quency.
Conversely, the impedance of a capacitor has no real compo-
nent. Its imaginary component is a function of both capaci-
tance and frequency. The current through a capacitor is al-
ways 90 degrees out of phase with the voltage across it, with
current leading the voltage. Because the impedance of a ca-
pacitor varies inversely with frequency, at high frequencies a
capacitor acts as a short circuit - it's impedance tends toward
zero. At low frequencies (approaching dc) a capacitor acts as
an open circuit, and the impedance tends toward infinite.
The third simple electrical component is the inductor. Like a
capacitor, the current through an inductor is always 90 de-
grees out of phase with the voltage drop across it. However,
the phase shift is in the opposite direction - the current lags
behind the voltage. Also, as the frequency increases, the i m-
pedance of an i nductor increases. It acts as a short circuit at
low frequencies and as a large impedance at high frequen-
cies.
To determine the total impedance of a combination of
simple elements, you combine the impedance values of
the individual components according to some simple
rules. For two circuit elements in series, the combined
impedance is simply the vector sum of the individual
impedance values.
Z
S
= Z
1
+ Z
2
(10)
In complex number representation, the real parts must
be added together to form the real component of the

Circuit Element Impedance Equation

Z = R + 0 j j = -1


Z = 0 j / C = 2f


Z = 0 + j L = 2f



R jCR
2

Z = -
1+
2
C
2
R
2
1+
2
C
2
R
2


series combination and the imaginary parts must be added to
form the imaginary component of the combination.
Z
S
' + j Z
S
" = (Z
1
'+Z
2
') + j (Z
1
+Z
2
") (11)
For example, if you have a resistor and capacitor in
series, the impedance expression will be the sum of the
impedance of the resistor (which has only a real part -
the imaginary part is identically zero) and the imped-
ance of the capacitor (which has only an imaginary part
- the real part is zero). This is shown in Table 2.
For the more complex parallel resistor/capacitor net-
work, the impedance expression becomes more com-
plicated. For circuit elements in parallel, the admittance
values (admittance being the inverse of impedance)
must be added together. Thus, for two impedance val -
ues in parallel,
1 1 1
---- = ----- + ------
Z
p
Z
l
Z
2
(12)

You can build more complicated equivalent circuits from
a series or parallel combination of simpler sub-circuits.
Plot Analysis
You can study an equivalent circuit by deriving its impedance
equation. However, its simpler to perform a measurement on
the circuit and analyze the resulting plot. Youll get a good
picture of the real and imaginary impedance components and
of the phase shift characteristics as a function of frequency.
The Randles cell (see Figure 5) models the electrochemical
impedance of an interface and fits many chemical systems.
You can easily equate the circuit components in the Randles
cell with familiar physical phenomena, such as adsorption or
film formation.
R is the ohmic or uncompensated resistance of the solution
between the working and reference electrodes.
Polarization Resistance versus
Charge Transfer Resistance
You can define resistance as the slope of the steady -sate poten-
tial vs. current plot. As such, it is the low frequency 9dc) limit for
the real part of the impedance. You can determine the polariza-
tion resistance by extrapolating the Nyquist plot to the real (x)
axis as the frequency approaches 0 Hz (dc).
On the other and, the first intersection of the Nyquist plot with
the real axis indicates charge transfer resistance. IN the simple
cases, the polarization resistance and the charge transfer resis-
tances are identical. However, for more complicated cases they
may not be equal (see Figure 8).
For both types of resistance, you must subtract the solution
resistance to get an accurate value.
R
p
is the polarization resistance or charge-transfer re-
sistance at the electrode/solution interface. C
DL
is the
double layer capacitance at this interface.
If you know the polarization or charge-transfer resis-
tance, you can calculate the electrochemical reaction
rates. Double-layer capacitance measurements can
provide information on adsorption and desorption phe-
nomena. In some systems, a C
DL
measurement may
not represent the double layer capacitance. Rather, it
may indicate the degree of film formation or the integrity
of an organic coating.
39

The impedance of a capacitor diminishes as the fre-
quency increases (see TabIe 2), while the impedance
of a resistor is constant. Thus, above a certain fre-
quency, the impedance of the capacitor, C
DL
, becomes
much smaller than the impedance of the resistor, R
p
.
Since C
DL
is in parallel with R
p
, the capacitor acts as a
short and effectively removes the resistor from the cir-
cuit. At the highest frequencies, the impedance of the
capacitor will also become much smaller than R

. Thus,
the high frequency behavior of the Randles cell is con-
trolled almost entirely by R

.
However, at the lowest frequencies, the capacitor acts
as an open circuit and is effectively removed from the
circuit. The impedance of the Randles cell is then the
combined resistance values of the two series resistors
R

and R
p
.
Thus, at both the high and the low frequency limits, the
Randles cell behaves primarily as a resistor. The
imaginary component is very small, the phase angle is
close to 0 degrees, and the impedance does not
change with frequency. At intermediate frequencies, the
capacitor's impedance begins to have an effect and the
cell becomes more capacitive. The imaginary compo-
nent becomes significant, the phase angle can start to
approach 90 degrees, and the cell impedance becomes
frequency dependent.
Figure 6 shows an equivalent circuit proposed for a cor-
roding metal coated with a porous, non-conductive
film.
40
The additional circuit elements are the coating
capacitance (C
c
) and the pore resistance (R
po
).
The circuit in Figure 7 represents an electrochemical
reaction coupled to a chemical reaction. Here, R
CR
and
C
CR
represent the resistive and capacitive effects of the
chemical reaction. The boxed sub-circuit is a simplified
illustration of the effects of diffusion. The enclosed re-
sistor and capacitor, R
W
and C
W
, provide a rough ap-
proximation of the Warburg impedance, a value used to
account for mass transfer limitations due to diffusion
processes adjacent to the electrode.
To determine which equivalent circuit best describes
the behavior of an electrochemical system, you must
measure the impedance over a range of frequencies.
The standard technique is to apply an ac voltage or
current over a wide range of frequencies and measure
the current or voltage response of the electrochemical
system. You can then calculate the system's imped-
ance by analyzing the response signal at each fre-
quency.
To completely describe the behavior of an electro-
chemical system, you must know the values of both the
in- phase and out -of-phase impedance components at
a number of frequencies across the range of interest.
You can calculate these values by applying Equation 5
to the real and imaginary components of the excitation
and response waveforms. You can characterize most
electrochemical systems quite well by gathering imped-
ance data in the 0.001 Hz to 1 x IO
4
Hz frequency
range.
R


C
DL

R
P

R = uncompensated resistance
RP = polarization resistance
CDL = double layer capacitance
C
DL

R
P

R


C
C

R
C

RCR = Chemical Reaction Resistance
CCR = Chemical Reaction Resistance
RW & CW = Warburg Impedance Effects
Figure 5: Equivalent Circuit for a Single Electrochemical Cell
Figure 6: Equivalent Circuit for a Metal Coated with a Porous,
Non-conductive Film
Figure 7: Equivalent Circuit for an Electrochemical Reaction
Coupled to a Chemical Reaction
R


C
W
R
W
C
CR
R
CR
R
t

C
DL


Overview
Once an experiment is complete, the raw data at each
measured frequency consists of these components:
The real component of voltage (E)
The imaginary component of voltage (E)
The real component of current (I)
The imaginary component of current (I)
From this raw data you can compute the phase shift
( ) and the total impedance (Z) for each applied fre-
quency, as well as many other impedance functions.
You can use a variety of formats to plot this data. Each
format offers specific advantages for revealing certain
characteristics of a given chemical system. You can
discover the true behavior of a real chemical system
only by looking at all of the available plotting formats.

The Nyquist Plot
Figure 8 shows one popular format for evaluating elec-
trochemical impedance data, the Nyquist plot. This for-
mat is also known as a Cole-Cole plot or a complex
impedance plane plot. In our study, we plotted the
imaginary impedance component (Z) against the real
impedance component (Z') at each excitation fre-
quency. The plot in Figure 8 illustrates the expected
response of the simple circuit in Figure 5.
We saw that at high frequencies, the impedance of the
Randles cell was almost entirely created by the ohmic
resistance, R

. The frequency reaches its high limit at


the leftmost end of the semicircle, where the semicircle
touches the x axis. At the low frequency limit, the Ran-
dies cell also approximates a pure resistance, but now
the value is (% + Rp). The frequency reaches its low
limit at the rightmost end of the semicircle.
The Nyquist plot has several advantages. The primary
one is that the plot format makes it easy to see the ef-
fects of the ohmic resistance. If you take data at suffi-
ciently high frequencies, it is easy to extrapolate the
semicircle toward the left, down to the x axis to read the
ohmic resistance. The shape of the curve (often a semi-
circle) does not change when the ohmic resistance
changes. Consequently, it is possible to compare the
results of two separate experiments that differ only in
the position of the reference electrode! Another advan-
tage of this plot format is that it emphasizes circuit
components that are in series, such as R

.
The Nyquist plot format also has some disadvantages.
For example, frequency does not appear explicitly.
Secondly, although the ohmic resistance and polariza-
tion resistance can be easily read directly from the Ny-
quist plot, the electrode capacitance can be calculated
only after the frequency information is known. As shown
in Figure 8, the frequency corresponding to the top of
the semicircle,
( = MAX)
, can be used to calculate the
capacitance if R
p
is known.
Although the Nyquist format emphasizes series circuit
elements, if high and low impedance networks are in
series, you will probably not see the low impedance
circuit, since the larger impedance controls plot scaling.
Figure 8 illustrates this point.
The Bode Plot
Figure 9 shows a Bode Plot for the same data pictured
in the Nyquist plot in Figure 8. The Bode plot format lets
you examine the absolute impedance, |Z|, as calculated
by Equation 8, and the phase shift, , of the impedance,
each as a function of frequency.
The Bode plot has some distinct advantages over the
Nyquist plot. Since frequency appears as one of the
axes, it's easy to understand from the plot how the im-
pedance depends on the frequency. The plot uses the
logarithm of frequency to allow a very wide frequency
range to be plotted on one graph, but with each decade
given equal weight. The Bode plot also shows the mag-
nitude ( | Z | ) on a log axis so that you can easily plot
wide impedance ranges on the same set of axes. This
can be an advantage when the impedance depends
strongly on the frequency, as is the case with a capaci-
tor.
The log I Z I vs. log curve can yield values of R
p
and
R

. At the highest frequencies shown in Figure 9, the


ohmic resistance dominates the impedance and log
(R

) can be read from the high frequency horizontal


plateau. At the
Impedance Plots
0
10
20
30
40
50
60
70
80
0 10 20 30 40 50 60 70 80 90 100 110 120
R+ R
P
R RP = 2 /Z/ tan

max

max Z
= 1/
CRP
, = 2f


Figure 8: Nyquist Plot for a Simple Electrochemical System
Decreasing frequency
I
m
a
g
i
n
a
r
y

Real

max Z

lowest frequencies, polarization resistance also contri b-
utes, and log (R

+ R
p
) can be read from the low fre-
quency horizontal plateau. At intermediate frequencies,
this curve should be a straight line with a slope of -1.
Extrapolating this line to the log I Z I axis at = 1 (log
= 0, f = 0.16 Hz) yields the value of C
DL
from the rela-
tionship:
|Z| = 1/C
DL
(I 0)
where = 2f
The Bode plot format also shows the phase angle, . At
the high and low frequency limits, where the behavior of
the Randles cell is resistor-like, the phase angle is
nearly zero. At intermediate frequencies, increases as
the imaginary component of the impedance increases.
The vs. log plot yields a peak at
( =MAX)
, the fre-
quency, in radians, at which the phase shift of the re-
sponse is maximum. The double-layer capacitance,
C
DL
, can be calculated from Equation 11.

( =MAX)
= (1 / C
DL
R
p
) (1 + R
p
/R) (11)
Note that both R
p
and R

appear in this equation! It is


important to remember that this frequency will not be
the same as the frequency at which the Nyquist plot
reaches its maximum.
The Bode plot is a useful alternative to the Nyquist plot.
It lets you avoid the longer measurement times associ-
ated with low frequency R
p
determinations. Further-
more, the log |Z| vs. log plot sometimes allows a
more effective extrapolation of data from higher fre-
quencies.
The Bode format is also desirable when data scatter
prevents adequate fitting of the Nyquist semicircle. In
general, the Bode plot provides a clearer description of
the electrochemical system's frequency-dependent be-
havior than does the Nyquist plot, in which frequency
values are implicit rather than explicit.
On some electrochemical processes, there is more than
one rate-determining step. Each step represents a sys-
tem impedance component and contributes to the over-
all reaction rate constant. The electrochemical imped-
ance experiment can often distinguish among these
steps and provide information on their respective rates
or relaxation times.
Figure 10 is typical of multiple time-constant Nyquist
plots. While close inspection reveals two semicircles,
one of the semi -circles is much smaller than the other,
making it difficult to recognize multiple time constants.

Log = 0 = 0
0
o
90
o /Z/=1/C
dl

max
R

R
p
+R

Figure 9: Bode Plot for a Simple Electrochemical System
Log |Z|

Evaluating Capacitance
from a Bode Plot
At intermediate frequencies, the impedance of the capacitor C,
can control the total impedance of the Randles cell. The capaci-
tor becomes the controlling component whenever RW and RP
differ by more than a factor of 100 or so. Under these conditions:
ZC = -j / C
And
log |Z| = log (1/ C)
= -log (C)
= -log () log (C)
= -log (2f) log (C)
= -log (2) log (f) log (C)
Note that the Bode plot of log |Z| vs log (f) or log () has a slope
of 1 in this region.
At this point f = 0.16 Hz, = 2f = 1 and log (2f) = 0
Therefore:
Log (|Z|) = -log (C) (when f = 0.16 Hz)
or
|Z (f = 0.16 Hz) | = 1/C


Figure 10: Nyquist Plot for a Two Time Constant Cell
0
10000
20000
30000
40000
50000
60000
70000
0 20000 40000 60000 80000 100000
Z
im

(
o
h
m
s
)
Zre (ohms)
Figures 11 and 12 show Bode plots for the same data
shown in Figure 10. The Bode plot format lets you eas-
ily identify the frequency break points associated with
each limiting step.
The Bode plot also has some disadvantages. The great
est one is that the shape of the curves can change if
the circuit values change. Figures 13 and 14 show
Bode plots for several similar circuits. The only differ-
ences are the values of the uncompensated resistance.
Note that the location (= MAX) and height of the phase
maximum depend on the value of R

. Also note that the


slope of the central portion of the log |Z| plot is


Influenced by the value of R

as well. The R

value can
have an effect on the capacitance values calculated
from Equation 1. The corresponding Nyquist plots all
have the same semicircle shape (see Figure 15).
The Randles Plot
The Randles plot is useful in determining whether War-
burg impedance is a significant component of the
equivalent circuit model. Identifying the presence of
Warburg impedance can help you describe a reaction
mechanism. A slope of -1/2 or -1/4 in the linear portion
of the Bode plot can also indicate diffusion control.
Figure 16 shows an idealized Randles plot of Z' vs.
for a diffusion-controlled system. In this case Z' and Z
are equal and are linear functions in .
For a completely reversible system under pure diffusion
control, the mass transfer (Warburg) impedance, Z
w
, is
given by:
Z
W
= s 2
(12)
where s is the Warburg coefficient, from which the di f-
fusion coefficient may be calculated.
Figure 11: Bode Plot for a Two Time Constant Cell (Impedance vs.
Frequency)
Figure 12: Bode Plot for a Two Time Constant Cell
(Phase Angel vs. Frequency)
0
10
20
30
40
50
60
70
80
90
0.00001 0.01 10 10000 10000000
p
h
a
s
e

o
f

Z

(
d
e
g
)
Frequency (Hz)
1.0 E+00
1.0 E+01
1.0 E+02
1.0 E+03
1.0 E+04
1.0 E+05
1.0 E+06
0.00001 0.01 10 10000 10000000
|
Z
|

(
o
h
m
s
)
Frequency (Hz)
1 . 0 E - 0 1
1 . 0 E+ 0 0
1 . 0 E+ 0 1
1 . 0 E+ 0 2
1 . 0 E+ 0 3
1 . 0 E+ 0 4
1 1000
|
Z
|

(
o
h
m
s
)
Frequency (Hz)
1000
100
10
0
Figure 13: Bode Plots for Selected Values of R
(Impedance vs. Frequency)
Figure 15: Nyquist Plots for Different Values of R
Figure 14: Bode Plots for Selected Values of R (Phase Angle
vs. Frequency)
0
50
100
100 100000
p
h
a
s
e

o
f

Z

(
d
e
g
)
Frequency (Hz)
0
500
1000
0 500 1000 1500 2000
Z
i
m

(
o
h
m
s
)
Zre (ohms)
100
1000


1000
100
10
0
Thus the linearity of the Randles plot can be used as a
test of diffusion control, and in certain cases the War-
burg diffusion coefficient can also be calculated from
the slope.
Other Plot Formats
Other plot formats optimize data interpretation for spe-
cific experimental systems. Figure 17 shows how Ra,
Rp and CE)L are derived from a plot of Z' vs. wZ'. For
the Randles cell, this plot has the advantage of being a
simple straight line. Moreover, once you know the Rp
value, you can easily calculate the capacitance. This
format provides a more reasonable fit of scattered data
than the Nyquist plot.
You can sometimes get clearer picture of the system's
behavior from a plot of Y'/ vs Y'/ , also known as a
capacitance plot (see Figure 18). The admittance, Y, is
simply the inverse of impedance.

Y = I = 1
E Z (13)
It's easy to see from Equation 13 that the admittance of
two circuit elements in parallel is simply the sum of the
individual admittances. For this reason, plots involving
admittance often emphasize circuit elements in parallel.
The impedance data shown in Figure 15 can easily be
converted to the format shown in Figure 18, which per-
mits a convenient determination of system capacitance
as Y / .
For a pure capacitor, Y
c
= j C, or Y
c
/ = j C. There-
fore, on this plot a capacitor will have only an imaginary
component, and (Y
c
/ ) will be independent of fre-
quency. On the capacitance plot, a capacitor will be
represented by simply a dot on the y axis.
For a resistor, Y
R
= (1/R) and Y
R
/ = (1/R) / . In con-
trast to the capacitor, the resistor will have only a real
component, but its value will depend on the frequency.
A resistor will appear as a horizontal straight line on the
capacitance plot of Figure 18.





Warburg Impedance
The rate of an electrochemical reaction can be strongly influ-
enced by diffusion of a reactance towards or a product away
from the electrode surface. This is often the case when a
solution species must diffuse through a film on the electrode
surface. This situation can exist when the electrode is covered
with reaction products, adsorbed solution components, or a
prepared coating. Whenever diffusion effects completely
dominate the electrochemical reaction mechanism, the imped-
ance is called the Warburg Impedance.
For diffusion-controlled electrochemical reaction, the current is
45 degrees out of phase with the imposed potential.
41
With this
phase relationship, the real and imaginary components of the
impedance vector are equal at all frequencies. In terms of
simple equivalent circuits, the behavior of Warburg impedance
(a 45 degree phase shift) is midway between that of a resistor
(a 0 degree phase shift) and a capacitor (90 degree phase
shift). There is no simple electrical equivalent for the Warburg
impedance.
Figure 16: Idealized Randles Plot of Z" vs.
R


Z (ohms)
W


(rad/sec)
1/2
Y/ (Farad)
Y


(
F
a
r
a
d
)
1.000E-4
1000
250
10
50
0
2.000E-4
Figure 17: A Plot of Z' vs. Z"
Z
Z
1/C
R
Z = R + R
P
R
P
CZ
R
P
= R
W

Figure 18: A Capacitance Plot
When a resistor and capacitor are in parallel, the ca-
pacitance plot is simply the sum of the impedance for
the resistor and the capacitor. This will give the horizon-
tal line shown in Figure 18. The semicircle shape is a
result of the interaction of the ohmic resistance and the
capacitance. The value of the capacitance can be read
from the extrapolation of the semicircle to the y-axis.
Under some conditions it may also be read from the
Y'/w value on the horizontal line section of the plot. This
is only possible if the point formed by the line and semi -
circle is a sharp one, as illustrated in Figure 18.
Data Interpretation
Although the simple equivalent-circuit experiments
generate relatively straightforward results, typical elec-
trochemical analyses yield more complicated plots.
These complexities arise because the simple equivalent
circuits do not fully describe the physical phenomena of
an electrochemical system. Yet simple equivalent circuit
models are frequently good approximations of real sys-
tems, and data can often be fitted to yield results of
reasonable accuracy.
Several computer programs have been written to fit
experimental data to a simple equivalent circuit model.
Because of the complexity of this problem, all of these
programs require some initial guesses for the circuit
parameters.
Figure 19 shows a Nyquist plot for an iron specimen in
deaerated 1N H
2
SO
4
. Note that the plot is not a perfect
semicircle, having an additional loop at low frequencies.

Figure 19: Nyquist Plot of Iron in Deaerated 0.1 N H2SO4 after
4 Hour Immersion.
This distortion has been attributed to the inductive be-
havior of the electrochemical system .
42,43

Several analytical treatments for such a plot have been
proposed. Most investigators believe that it is incorrect
to interpret either of the actual low frequency Z' axis
intercepts as the sum of P

and R
p
. Instead, they use
various curve-fitting techniques to obtain these values.
The Nyquist plot in Figure 19 reveals another complex-
ity. In this plot, the center of the circle does not lie on
the x-axis, but below it. To explain this depressed semi -
circle phenomenon, some researchers use models that
assume that the surface of the electrode is not homo-
geneous.
44
Using this model, you can characterize di f-
ferent areas of the surface with different time constants.
In this case, the total impedance of the surface would
be the parallel combination of these areas, You can use
several RC sub-circuits in parallel to model the imped-
ance.




You can derive the complete impedance equation for
the simple Randles cell. However, since a Randles cell
contains resistors and capacitors in parallel, it's easier
to express the impedance values in terms of admit-
tance, the reciprocal of impedance. Although the sym-
bol for admittance is Y, we'll use 1/Z in this discussion.
The Randles cell is defined as a parallel combination of
capacitance and polarization or charge transfer resis-
tance. To calculate the total admittance of the cell, you
must first add the individual admittance values of the
parallel resistors and capacitors:
1
=
1
+
1
Z Z
R
Z
C


Polarization Resistance
The impedance expression for polarization or charge
transfer resistance (either or both symbolized here by
R
p
) is simple:
Z
R
= (R
p
+ 0j) = R
p

The admittance expression is:
1
=
1
=
1
Z
R
R
p
+ 0j R
p


Capacitance
The impedance expression for the capacitance is more
complex:
Z
C
= -j ( 1/ C )
The admittance expression is:
1
=
1
=
1
Z
c
: (0 j [ 1 / C] ) (-j [1/ C ] )
or
=
C
=
j C
=
j C
j (j) (-j) (- j
2
)
or finally, remembering that j
2
= -1:
=
j C
=
j C
-(-I)
Parallel Combination (R
p
and C)
You can calculate the admittance of the parallel combi-
nation from:
1/Z = 1/ZR + 1/Z
C
or
Z = 1/R
P
+ (jC) = 1/R
P
+ j(C)
or finally, after multiplying both top and bottom by R
P
:


1 R
P


(1/R
P
) + ( j [C] ) 1 + j(R
P
C)

For those unfamiliar with complex numbers:
1 1 a-jb a-jb
a+jb a+jb a-jb a
2
+ b
2

Thus, the impedance expression can be rewritten as:
R
P
j (R
P
2
C)
1 + (R
P
C)
2

Ohmic Resistance
Finally, the impedance of the ohmic resistance must be
added:
R
P
j (R
P
2
C)
1 + (R
P
C)
2

or
R
P
-(R
P
2
C)
1 + (R
P
C)
2
1 + (R
P
C)
2

or
= Z + j Z
Although this equation is correct, its complexity ob-
scures its meaning. It is useful, however, to check the
behavior at the two frequency limits, w = 0 and w = co.
As w tends towards zero, the denominator in both terms
tends toward 1. For the real term, the sum approaches
(R

+ R
P
) as goes to zero. For the imaginary term,
the numerator tends toward zero, so the imaginary term
vanishes at the dc limit.
As the frequency increases, this behavior may not be
as easy to see. In the real term, as w increases, the de-
nominator approaches (R
P
C)
2
since that term be-
comes much larger than 1. Consequently, the second
term ratio approaches zero and the sum (Z') becomes
simply R

. In the imaginary term, the ratio approaches


-(R
P
2
C) -1
(R
P
C)
2
C
which tends towards zero as w increases. Thus, at the
low frequency limit, the imaginary term vanishes and
the real part approaches (R

+ R
P
). At the high fre-
quency limit, the imaginary term also vanishes, but the
real term approaches R

.
Appendix: Deriving the
Randles Cell Impedance
) ( ( )
=
=
= = Z
Z =
= Z
+
(
)
j +
=
R


=
1. Mansfeld, F. "Recording and Analysis of AC Impedance Data for
Corrosion Studies: 1. Background and Methods for Analysis" Corro-
sion 1981, 37, 301-307.

2. Mansfeld, F.; Kendig, M. W.; Tsai, S. "Recording and Analysis of AC
Impedance Data for Corrosion Studies: 1. Experimental Ap- proach
and Results" Corrosion 1982, 38, 570-580.

3. Macdonald, D. D.; Syrett, B. C.; Wing, S. S. "Me Corrosion of Copper-
Nickel AJloys 706 and 715 in Flowing Seawater: 1. Effect of Oxygen"
Corrosion 1978, 34, 289-30 1.

4. Macdonald, D. D.; Syrett, B. C.; Wing, S. S. "The Corrosion of Cop-
per-Nickel Alloys 706 and 715 in Flowing Seawater: II. Effect of Dis-
solved Sulfide" Corrosion 1979, 35, 367-378.

5. Lorenz, W. J.; Mansfeld, F. "Determination of Corrosion Rates by
Electrochemical DC and AC Methods" Corrosion Science 1981, 21,
647.

6. Feigenbaum, C.; Gal-Or, L; Yahalom, J. "Scale Protection Criteria in
Natural Waters" Corrosion 1978, 34, 133-137.

7. Eppelboin, I.; Keddam, M.; Takenouti, H. "Use of Impedance Meas-
urements for the Determination of the Instant Rate of Metal Corrosion"
J. Appi. Electrochem. 1 972, 2, 7 1.

8. Keddam, M.; Mattos, 0. R.; Takenoud, H. "Reaction Model for Iron
Dissolution Studied by Electrode Impedance" (I and 11) J. Electro-
chem. Soc. 1981, 128, 257-274.

9. Cahan, B. D.; Chen, C. T. "Me Nature of the Passive Film on Iron: II.
Impedance Studies" J. Electrochem. Soc. 1982, 129, 474-480.

10. Glarum, S. H.; Marshall, J. H. "An A-C Admittance Study of the Plat i-
num/Sulfuric Acid Interface" J. Electrochem. Soc. 1979, 126, 424-430.

11. Isaacs, H. S.; Olmer, L J. "Correlation of the AC and DC Polarization
Resistances of a Platinum Electrode/Zirconia Solid Oxide Electrolyte
Interface" J. Electroanal. Chem. 1982, 132, 59-65.

12. Glarum, S. H.; Marshall, J. H. "The A-C Response of Nickel Oxide
Electrode Films" J. Electrochem. Soc. 1982, 129, 535- 542.

13. Bonnel, A. et al "Corrosion Study of a Carbon Steel irk Neutral Chlo-
ride Solutions by Impedance Techniques" J. Electrochem. Soc. 1983,
130, 753-761.

14. Dabosi, F. et al "Corrosion Inhibition Study of a Carbon Steel in Neu-
tral Chloride Solutions by Impedance Techniques" J. Electrochem.
Soc. 1983, M, 761-766.

15. Treatise on Materials Science and Technology, Vol. 23, Corrosion:
Aqueous Processes and Passive Films; Scully, J. C. Ed.; Academic
Press, Inc.: London, 1983.

16. Mansfeld, F.; Kendig, M. W.; Tsai, S. "Evaluation of Corrosion Behav-
ior of Coated Metals with AC Impedance Measurements", Corrosion
1982, 38, 478-485.

17. Padget, J. C.; Moreland, P. J. "Use of AC Impedance in the Study of
the Anticorrosive Properties of Chlorine-Containing Vinyl Acrylic La-
tex Copolymers" J. Coatings Tech. 1983, 55.

18. Walter, G. W. "Application of Impedance Measurements to Study
Performance of Painted Metals in Aggressive Solutions" J. Electro-
anal. Chem. 1981, 118, 259-273.
19. Standish, J. V.; Leidheiser, Jr., H. 'The Electrical Properties of Organic
Coatings on a Local Scale-Relationship to Corrosion" Corrosion 1980,
36, 390-395.

20. Scantlebury, J. D. et al "Simulated Underfilm Corrosion of Coated
Mild Steel Using an Artificial Blister" Corrosion 1983, 39, 108-112.

21. Leldheiser, Jr., H. "Towards a Better Understanding of Corrosion Be-
neath Organic Coatings" Corrosion 1983, 39, 189-201.

22. Sluyters, J. H. "On the Impedance of Galvanic Cells: 1. Theory' Re-
cueil 1960, 79, 1092.

23. Sluyters, J. H.; Oomen, J. J. C. "On the Impedance of Galvanic Cells:
11. Experimental Verification" Recueil 1960, 79, 1 101.

24. Sluyters, J. H.; Rehbach, M. "On the Impedance of Galvanic Cells: Ill.
Applications to Alternating Current Polarography Recueil 1961, 80,
469.

25. Sluyters, J. H.; Rehbach, M. "On the Impedance of Galvanic Cells: IV.
Determinations of Rate Constants of Rapid Electrode Reactions from
Electrode Impedance Measurements" Recueil 1962, 81, 301.

26. Zimmerman, A. H. et al "Impedance and Mass Transport Kinetics of
Nickel Cadmium Cells" J. Electrochem. Sac. 1982, 129, 289-293.

27. Casson, P.; Hampson, N. A.; Willors, M. J. "Fundamentals of Lead-
Acid Cells: Part VII. The A. C. Response of Lead Dioxide Electrodes
in Sulfuric Acid" J. Electroanal. Chern. 1979, 97, 21- 32.

28. McBreen, J. et al "Zinc Electrode Morphology in Alkaline Solutions:
11. Study of Alternating Charging Current Modulation on Pasted Zinc
Battery Electrodes" J. Electrochem. Soc. 1983, 130, 1641-1645.

29. Epelboin, I.; Joussellin, M.; WiarL R. "Impedance Measurements for
Nickel Deposibon in Sulfate and Chloride Electrolytes" J. Appi.
ElecLrochem. 1981, 119, 61.

30. Glarum, S. H.; Marshall, J. H. "An Admittance Study of the Copper
Electrode" J. Electrochem. Soc. 1981, 128, 968-979.

31. Baranski, A.; Fawcett, W. R. "Medium Effects in the Electroreduction
of Alkali Metal Cations" J. Electroanal. Chem. 1978, 94, 237-240.

32. Kisza, A.; Grzeszczuk, M. "Electrode Processes in Fused Organic Salts
Studied by Means of Impedance Method: Part 1. The Cadmium Elec-
trode in Fused Organic Salts" J. Electroanal. Chem. 1978, 91, 115-125.

33. Franceschetti, D. R.; Macdonald, J. R. 'Small-Signal A-C Response
Theory for Electrochromic Thin Films" J. Electrochem. Soc. 1982,
129, 1754-1756.

34. McCann, J. F.; Badwal, S. P. S. "Equivalent Circuit Analysis of the
Impedance Response of Semiconductor/Electrolyte/Counter-electrode
Cells" J. Electrochem. Soc. 1982, 129, 551-559.

35. Etman, M.; Koehler, C.; Parsons, R. "A Pulse Method for the Study of
the Semiconductor-Electrolyte Interface" J. Electroanal. Chem. 1981,
130, 57-67.

36. Weber, M. F.; Schumacher, L. C.; Dignam, M. J. "Effect of Hydrogen
on the Dielectric and Photoelectrochemical Properties of Sputtered
TiO2 Films" J. Electrochem. Soc. 1982, 129, 2022. 2028.


37. Macdonald, D. D. Transient Techniques in Electrochemistry; Plenum:
New York, 1977.

38. Macdonald, J. R. Impedance Spectroscopy: Ernphasizing Solid Mate-
rials and Systems; John Wiley & Sons: New York, 1987.

39. Padget; Moreland.

40. Cahan, B. D.; Chen, C. T. "Questions on the Kinetics of 02 Evolution
on Oxide-Covered Metals" J. Electrochem. Soc. 1982, 129, 700-705.
41. Macdonald, J. R., p 23.

42. Gabrielli, C. Identification of Electrochemical Processes by Frequency
Response Analysis (Monograph); Solartron Instrumentation Group
1980; pp 53-61.

43. Epelboin, 1. et a] "A Model of the Anodic Behaviour of Iron in Sul-
phoric Acid Medium" Electrochim. Acta 1975, 20, 913-916.

44. Lorenz; Mansfeld.

S-ar putea să vă placă și